Vous êtes sur la page 1sur 11

Journal of

Electroanalytical
Chemistry
Journal of Electroanalytical Chemistry 566 (2004) 111121
www.elsevier.com/locate/jelechem

Experimental and theoretical study of


1-(2-ethylamino)-2-methylimidazoline as an inhibitor of carbon
steel corrosion in acid media
J. Cruz a, R. Martnez a, J. Genesca b, E. Garca-Ochoa
a

a,*

Programa de Ingeniera Molecular, Instituto Mexicano del Petroleo, Eje Central Lazaro Cardenas No. 152, San Bartolo Atepehuacan,
C.P. 07730 Mexico DF
b
Departamento Ingenieria Metalurgica, Facultad Quimica, UNAM, Ciudad Universitaria, 04510 Mexico DF
Received 7 March 2003; received in revised form 21 August 2003; accepted 14 November 2003

Abstract
The electrochemical behavior of 1-(2-ethylamino)-2-methylimidazoline (imidazoline), its precursor N-[3-(2-amino-ethylaminoethyl)]-acetamide (amide) and its derivative 1-(2-ethylamino)-2-methylimidazolidine (imidazolidine), is evaluated by using potentiodynamic polarization curves and electrochemical impedance spectroscopy, EIS, techniques in deaerated acid media to compare
their corrosion inhibition eciency. The experimental results suggest that imidazoline is a good corrosion inhibitor at dierent
concentrations whereas amide shows low eciency values; however, the properties of a corrosion inhibitor were not found in
imidazolidine. The reactivity of these compounds was analyzed through theoretical calculations based on density functional theory
(DFT) to explain the dierent eciencies of these compounds as corrosion inhibitors both in the neutral and protonated form. The
theoretical results indicate that imidazoline is the more ecient corrosion inhibitor because of its two very active sites (two nitrogen
atoms) and the plane geometry of the heterocyclic ring, thus promoting coordination with the metal surface.
2003 Elsevier B.V. All rights reserved.
Keywords: Amide; Corrosion inhibitors; Imidazoline; Imidazolidine; EIS; DFT calculations

1. Introduction
Nowadays the study of carbon steel corrosion phenomena has become an important industrial and academic topic [1]. The use of corrosion inhibitors is one of
the most eective methods to protect metal surfaces
against corrosion, especially in acid media [2,3].
The development of corrosion inhibitors is based on
organic compounds containing nitrogen, oxygen, sulfur
atoms, and multiple bonds in the molecules that facilitate adsorption on the metal surface [4]. The corrosion
inhibition eciency of organic compounds is related to
their adsorption properties. Adsorption depends on the
nature and the state of the metal surface, on the type of
corrosive medium and on the chemical structure of the
inhibitor [4]. Studies report that the adsorption of the
*

Corresponding author.
E-mail address: eochoa@imp.mx (E. Garca-Ochoa).

0022-0728/$ - see front matter 2003 Elsevier B.V. All rights reserved.
doi:10.1016/j.jelechem.2003.11.018

organic inhibitors mainly depends on some physicochemical properties of the molecule, related to its functional groups, to the possible steric eects and electronic
density of donor atoms; adsorption is suppose also to
depend on the possible interaction of p-orbitals of the
inhibitor with d-orbitals of the surface atoms, which
induce greater adsorption of the inhibitor molecules
onto the surface of carbon steel, leading to the formation of a corrosion protecting lm [1,5].
Dierent derivatives from imidazoline are employed
as steel corrosion inhibitors. Even though they have
been specially employed in the oil industry, only recently
have many studies been undertaken to understand how
they work [6]. The availability of sophisticated computational tools and electrochemical techniques and a
better understanding of corrosion inhibitors and their
mechanisms can now be achieved [79]. Ramachandran
et al. [10], Wang et al. [11], and Cruz et al. [12] have
published important papers concerning the molecular

112

J. Cruz et al. / Journal of Electroanalytical Chemistry 566 (2004) 111121

structure of imidazoline as a corrosion inhibitor. In


these implies some key questions regarding the structureperformance relationships of imidazolines and
corresponding amines are: (a) the role of the hydrocarbon chain relative to the imidazoline head group and
pendant amine group in lm formation; (b) the thickness of the imidazoline lm; (c) the stability of the imidazoline lm; (d) the solution composition and
hydrolysis of imidazoline.
One of the most common methods used to prepare 2alkyl-imidazoline consists of the set of reactions depicted in Scheme 1 below [1317]. The thermal reaction
between 1,2-diamines (1) and carboxylic acids (2) leads
to an intermediary amide (3), which then cycles to (4) at
high temperatures and low pressure. The imidazoline
cycle can be reduced to a saturated imidazolidinic-like
class (5), which can be obtained when 1 reacts with an
aldehyde [17].
According to some reports, a high-rate conversion
from imidazoline to amide (up to 80%) has been reported [18] within a period of 29 days under atmospheric conditions, because of the low stability of the
imidazoline cycle, which can be hydrolyzed to (3) [18].
Other authors [19,20] have found that amide and
imidazoline have the same eciency as corrosion
inhibitors.
The purpose of this paper is to evaluate the behavior
of 1-(2-ethylamino)-2-methylimidazoline (imidazoline),
N-[3(2-amino-ethylamino-etyl)]-acetamide (amide) and
1-(2ethylamino)-2-methylimidazolidine (imidazolidine)
as corrosion inhibitors using potentiodynamic polarization curves and electrochemical impedance spectroscopy, EIS, techniques. The eect of the presence of the
iminic double bond N@C in the imidazoline ring and
the cyclic structure of imidazoline on its inhibition efciency is compared with those of amide and imidazolidine, in order to seek correlation between the
electronic and structural properties of the above compounds and their experimental behavior as corrosion
inhibitors. A theoretical study is carried out based on
density functional theory (DFT), which can provide
an interpretation of the experimental results obtained.
The structural and electronic parameters can be obtained by means of rst principles theoretical calculations using the computational methodologies of
quantum chemistry.

2. Experimental procedure
Electrochemical tests were carried out with an electrochemical interface Solartron 1287 connected to a
frequency response analyzer (FRA) Solartron 1255B,
which was controlled by CorrWare and Zplot software.
The cell assembly consisted of a carbon steel working
electrode (WE), two-graphite counter electrodes (CE),
and a saturated calomel electrode (SCE) as the reference
electrode. The test material was carbon steel (composition, wt%: 0.18 C, 0.35 Mn, 0.17 Si, 0.025 S, 0.03 P, and
bal. Fe). Prior to exposure, it was polished with 400-grit
silicon carbide (SiC) paper wetted with deionized water,
then polished with 600-grit SiC paper wetted with ethanol (C2 H6 O), and then rinsed with C2 H6 O. Twice
distilled water and analytical reagent-grade hydrochloric
acid, 0.5 M HCl, were used to prepare the test solutions,
which were de-aerated by means of nitrogen during a
period of 45 min. The samples were exposed at static
conditions and room temperature.
Polarization curves were recorded potentiodynamically at a rate of 0.5 mV/s in the range )300 to +300 mV
versus the open circuit potential. Impedance was measured over a frequency range of 100 kHz10 mHz. All
the impedance experiments were carried out after dipping the working electrode into 0.5 M HCl solution
containing the inhibitor at the open-circuit potential,
Ecorr , with respect to a SCE reference electrode, with a
10 mV peak-to-peak perturbation. Two independent
experiments have been recorded for each inhibitor
concentration value.
N-[3-(2-amino-ethylamino-ethy)]-acetamide (amide)
and 2-methyl-N-(2-etihylamino)-imidazoline (imidazoline) were synthesized as follows. A solution of 10 ml of
xylene in glacial acetic acid solution (6.0 g, 0.1 mol) was
slowly added to a diethylenetriamine solution previously
heated at 50 C (10.3 g, 0.1 mol) in a closed system
under a nitrogen atmosphere and constant agitation.
The addition having been completed, the solution was
heated until the azeotropic distillation of xylene/water
produced amide (93%). After this stage was completed,
the reaction was progressively heated to 190 C at reduced pressure (3 mm Hg) until all the water was distilled. The product obtained was puried in a
chromatographic column of silica-gel using petroleum
ether as an eluant, imidazoline (91%) was obtained.

Scheme 1.

J. Cruz et al. / Journal of Electroanalytical Chemistry 566 (2004) 111121

113

2-methyl-N-(2-ethylamino)-imidazolidine (imidazolidine): 10.3 g (0.1 mol) of diethylenetriamine, mixed with


11.0 g of acetaldehyde (solution at 40%) (0.1 ml) in 100
ml of tetrahydrofuran (THF), was agitated continuously
for 8 h at room temperature. The solvent was evaporated and the remaining product (raw) was puried in a
chromatographic column of silica-gel with hexane + ethyl acetate (9:1) as eluant, to give imidazolidine
in 79% yield.
The purity of the compounds was veried by means
of infrared spectroscopy, IR, 1 H and 13 C nuclear magnetic resonance (NMR), and mass spectroscopy (MS).

3. Computational details
Calculations were performed with the Gaussian 98
program [21]. Exchange and correlation calculations
were carried out with the functional hybrid B3LYP
[22,23] and the 6-311+G** orbital basis sets for all atoms. In all cases, total structure optimization together
with the vibrational analysis of the optimized structures
was carried out in order to determine whether they
corresponded to a maximum or a minimum in the potential energy curve. The frontier molecular orbitals,
namely, the highest occupied molecular orbital
(HOMO) and lowest unoccupied molecular orbital
(LUMO) and chemical reactivity were evaluated by
taking into account the Fukui indexes [24], which were
obtained by addition or subtraction of one electron,
N 1 and N  1.

4. Experimental results and discussion


The eect of the exposure time on the corrosion rate
was studied. The results obtained provided information
about the determination of the WE exposure time. The
corrosion current density, jcorr was recorded after an
exposure to 1, 2, 3, 4, 5, 6, 7, 8 and 9 h. The results are
presented in Fig. 1. It is seen that the corrosion rate
depends on the immersion time of carbon steel in the
solution. Fig. 1 illustrates the variation of jcorr with
immersion time. The corrosion rate increase initially and
then starts to decrease slowly. Thus it was found experimentally that a 9 h exposure seems necessary to
minimize of the eects of the electrodes preliminary
treatment and of the possible corrosion products
formed.
Fig. 1(a)(c) shows the estimated corrosion rate
during a 9 h period by means of the Tafel extrapolation
technique for the three organic compounds under study.
Fig. 1(a) corresponds to the corrosion rates obtained at
dierent amide concentrations showing some discrepancies in the behavior of the corrosion rate. From all the
inhibitor concentrations studied, 50 ppm show the

Fig. 1. Corrosion rate against time for carbon steel in 0.5 M HCl
containing dierent concentrations of: (a) amide; (b) imidazoline; (c)
imidazolidine.

higher corrosion rate, which is, however, less than the


corrosion rate corresponding to the blank solution, but
higher than the corresponding rate at 25 and 10 ppm.
One can conclude a non-monotonic dependence of the
corrosion rate on the concentration of amide.
Fig. 1(b) clearly shows a monotonic dependence of
the corrosion rate on the concentration of imidazoline.
In this case, the higher the imidazoline concentration,
the lower is the associated corrosion rate.

114

J. Cruz et al. / Journal of Electroanalytical Chemistry 566 (2004) 111121

Finally, Fig. 1(c), which corresponds to the corrosion


rate determined for the dierent imidazolidine concentrations, shows an erratic behavior. Again a nonmonotonic dependence of the corrosion rate on the
concentration of imidazolidine can be observed.
The Nyquist representation of EIS data obtained
before and after the addition of the three chemical
compounds at the dierent concentrations studied and
used as corrosion inhibitors after 9 h of being exposed in
the corrosive medium are provided in Fig. 2(a)(c). The
impedance spectra were similar, exhibiting in all cases a
single (capacitive like) semicircle and then only one time
constant, which can be modeled as an electric equivalent
circuit with a parallel combination of double-layer capacitance Cdl and polarization resistance Rp in series
with the solution resistance, Rs (Fig. 3), as has been reported previously for carbon steel studies in acid media
[2527]. With the addition of the water soluble chemicals, it is clearly seen that the diameter of the semicircle
increases with increasing inhibitor concentration.
However, the shape of the impedance spectra remained
unchanged showing a single time constant, which is the
result of the charge-transfer corrosion process.
It is important to point out that the three compounds
under study exhibited the same behavior during all test
times and for all concentrations. It can be observed that
as the inhibitor is adsorbed onto the metal surface, the
polarization resistance, Rp increases. As the inhibitor
adsorbs onto carbon steel, the capacitance of the interface starts decreasing. The change in the imaginary
component of the impedance and consequently of capacitance can thus be used to follow the adsorption
process. The double-layer capacitance Cdl reduces after
adsorption of the inhibitor since the adsorbed lm reduces the dielectric constant between the metal and
electrolyte. The large semicircles observed from high to
low frequencies, Fig. 2, indicates that the polarization
resistance, Rp becomes dominant due to adsorption of
the inhibitor.
In some cases impedance data obtained at the corrosion potential, Ecorr have the shape of depressed
semicircles with the center of the circle below the real
axis. Then the complex impedance Z (jx) could be expressed as
Z Rs

Rp
a:
1 jxCdl Rp

The a exponent is a unitless parameter that equals


one for an ideal capacitor. In most real systems, ideal
capacitive behavior is not observed due to surface
roughness, or other eects that causes uneven current
distributions on the electrode surface. In the case when
a 1, the term jxCdl Rp a reduces to jxCdl Rp , where Cdl
is the interfacial double-layer capacitance. As can be
appreciated in Table 1, in all the cases studied, a reaches
approximately the same value of 0.9. This can be in-

Fig. 2. Nyquist impedance plot for carbon steel in 0.5 M HCl containing dierent concentrations of: (a) amide; (b) imidazoline; (c) imidazolidine.

terpreted as an indication of the degree of inhomogeneity of the metal surface, corresponding to a small
depression of the double-layer capacitance semicircle,
and a has almost a constant value (0.9).
The impedance spectra of carbon steel in 0.5 M HCl
in the presence of the dierent inhibitors studied are

J. Cruz et al. / Journal of Electroanalytical Chemistry 566 (2004) 111121

Fig. 3. Equivalent circuit model for carbon steel in acid solution.

shown in Fig. 2(a)(c) in a Nyquist impedance plot. It is


found that all the inhibitors decrease the double-layer
capacitance and increase the polarization resistance of
the metaljelectrolyte interface to dierent extents, depending on the nature of the adsorption. As Rp value
increases, the inhibit power is higher.
A comparison of the inhibitor properties of amide,
imidazoline and imidazolidine indicates that imidazoline
is a much better inhibitor than amide and imidazolidine.
From Fig. 2 we can clearly observe that the larger value
of Rp corresponds to imidazoline, followed by amide
and imidazolidine, the Rp values for this last inhibitor
being very similar to those of the blank solution. These
observations suggest that imidazoline forms a much
better homogeneous lm compared to amide. It seems
necessary to obtain more information to understand
how the imidazoline adsorption proceeds. The latter
would be determined by the disposition of the molecule
on the metal surface in the adsorption step, depending
on its conguration and the number and disposition of
its active centers.

115

Fig. 4(a)(c) shows the behavior of Rp , as measured


from EIS. The plots correspond to inhibitor concentrations of 0, 10, 25, 50 and 100 ppm. The results show
clearly that the presence of imidazoline increases the Rp
values, Fig. 4(b), whereas for amide, Fig. 4(a), and especially imidazolidine, Fig. 4(c), a non-monotonic dependence of the Rp values on the inhibitor concentration
can be observed. These results conrm those previously
obtained by means of the Tafel extrapolation technique,
Fig. 1; thus, both techniques are consistent with the
observed behavior.
The corrosion rates and Rp values obtained from
potentiodynamic polarization curves and impedance
measurements as well as the calculated inhibition eciencies (IE) for dierent concentrations of these compounds are given in Table 1. It can be observed that the
values of corrosion inhibition eciency obtained are
similar, thus both techniques are consistent with the
observed behavior and seem adequate for these calculations. It is interesting to note that because the Tafel
extrapolation technique implies a larger perturbation of
the system under study than the impedance technique, it
seems more appropriate that the values are derived from
the faradic impedance technique.
The equations used to determine the inhibition eciency were:
%Ejcorr

jcorr  jcorr inh


 100;
jcorr

where %Ejcorr is the percentage eciency determined


by means of the Tafel extrapolation technique, jcorr is the
corrosion current density without inhibitor, and jcorr
(inh) is the same current density with inhibitor, and
%ERp

Rp inh  Rp
 100;
Rp inh

Table 1
Inhibition eciency for amide, imidazoline and imidazolidine
104 jcorr /A cm2

Ejcorr /%

Cdl /lF cm2

r/s

Rp =X cm2

ERp /%

[Amide]/ppm
0
10
25
50
100

7.41
4.08
2.56
5.41
2.17

0
45
65
27
70

2.74
1.93
1.81
7.23
1.61

0.913
0.917
0.917
0.924
0.917

0.40
0.47
0.59
1.95
0.67

23
38
51
42
65

0
39
54
45
64

[Imidazoline]/ppm
0
10
25
50
100

7.41
4.10
2.77
2.21
1.92

0
44
63
70
74

2.74
1.76
1.54
1.40
1.16

0.913
0.905
0.909
0.896
0.905

0.40
0.48
0.60
0.73
0.76

23
43
62
82
103

0
46
62
71
77

0
)48
0
13
12

2.74
3.27
3.02
3.12
2.38

0.913
0.885
0.891
0.885
0.894

0.40
0.39
0.39
0.48
0.21

23
19
20
24
29

0
)21
)15
4
20

[Imidazolidine]/ppm
0
7.41
10
11.0
25
7.40
50
6.48
100
6.54

116

J. Cruz et al. / Journal of Electroanalytical Chemistry 566 (2004) 111121

Fig. 4 that the Rp values are highest for imidazoline,


followed by amide and nally by imidazolidine, which
has Rp values of the same order of magnitude as those of
the acid solution without inhibitor.
From the impedance data, it is possible to obtain the
double-layer capacitance values, Cdl and its time constant, r, which are very important parameters for the
understanding of the mechanism of corrosion inhibition
for these substances. The frequency at the top of the
semicircle (when Z 00 is at maximum) is given by
xmax 2pfmax

1
:
Rp Cdl

The term on the right of the above equation dictates


the time scale of current response of the Randles circuit
of Fig. 3 to applied voltage. This time constant, in seconds, is dened for applied current as
r Rp Cdl
being then inversely proportional to the maximum value
of the imaginary term of the impedance:
00
r 2pf Z;max
1 :

The frequency at the top of the semicircle is not easy


to estimate from the phase angle Bode plot commonly
used, but can be obtained directly from the Z 00 vs. frequency Bode plot. Another method [28] involves an estimation of capacitance from the Bode plot of log jZj vs.
log frequency. If the solution resistance is small, as in
this case, the impedance modulus, at high frequency,
when xRC is large, is given by
log jZj  log x  log C:

Fig. 4. Polarization resistance, Rp , against time for carbon steel in 0.5


M HCl containing dierent concentrations of: (a) amide; (b) imidazoline; (c) imidazolidine.

where Rp is the polarization resistance without inhibitor,


and Rp (inh) is the polarization resistance with inhibitor.
The Rp values obtained from impedance measurements allow us to follow the corrosion inhibition
properties of the three compounds studied as a function
of time, as was previously done with the Tafel extrapolation technique (Fig. 1). From the SternGeary
equation, Rp has a value inversely proportional to the
corrosion current density, jcorr It can be observed from

At high frequencies the plot is a straight line, of slope


)1, and if extrapolated to x 1 then log jZj  log C
[28]. Fig. 5(a)(c) displays the capacitance values Cdl
obtained as a function of time for the three organic
compounds studied at dierent concentrations. The results indicate that the amide and imidazoline inhibitors
aect the values of the Cdl capacitance signicantly and
in all cases, show a clear tendency to increase as a
function of time. Fig. 5(a) showing the amide values,
indicates that the inhibitor concentration is inversely
proportional to the capacitance; however, for concentrations of 50 ppm, the capacitance value increases notably. Fig. 5(b) shows a monotonic dependence between
the imidazoline concentration and capacitance value,
where the imidazoline concentration is inversely proportional to the capacitance value. Finally, in Fig. 5(c)
the associated capacitance values of the imidazolidine
are very similar to those of the blank solution; this again
allows us to make conclusions about the small interaction between imidazolidine and the metal surface.
Fig. 6(a)(c) shows the variation of the time constant
r associated with the capacitor at dierent times as a
function of inhibitor concentration. While the capaci-

J. Cruz et al. / Journal of Electroanalytical Chemistry 566 (2004) 111121

Fig. 5. The double-layer capacitance against time for carbon steel in


0.5 M HCl containing dierent concentrations of: (a) amide; (b) imidazoline; (c) imidazolidine.

tance of a capacitor measures its capacity to store


charge, the time constant r is a measure of the rate at
which this charge can be stored. In Fig. 6(a) can be
observed the eect of amide concentration on r values
as a function of time. Again the r values corresponding
to 50 ppm show erratic behavior. In Fig. 6(b), which
corresponds to r values of imidazoline at dierent
concentrations, typical behavior is observed, the highest
values of r corresponding to a concentration of 100 ppm
of imidazoline and the lower to 10 ppm. The physical
meaning of this behavior takes into account that the

117

Fig. 6. The time constant against time for carbon steel in 0.5 M HCl
containing dierent concentrations of: (a) amide; (b) imidazoline; (c)
imidazolidine.

required time for the associated capacitor to be charged


or discharged is higher if the inhibitor concentration is
higher too. Fig. 6(c) shows the r values of imidazolidine.
In this case the values are similar, independently of the
inhibitor concentration, and of the same order of magnitude as for the solution without inhibitor, which
conrms what was indicated previously in the sense
that this substance does not aect the metal surface
signicantly.

118

J. Cruz et al. / Journal of Electroanalytical Chemistry 566 (2004) 111121

In general, signicant dierences can be appreciated


referring to the inhibitory eciency of imidazoline,
amide and imidazolidine. The understanding that the
corrosion inhibition eciency of organic compounds is
related to their adsorption properties allows us to propose a possible mechanism. The inhibition eciency
depends strongly on the structures and chemical properties of the species formed under the experimental
conditions studied. The extent of adsorption is dependent upon the electronic structure of the metal and the
inhibitor. The impedance spectrum of carbon steel in 0.5
M HCl in the presence of 100 ppm imidazoline gives a
polarization resistance of 103 X cm2 compared to the
blank value of 23 X cm2 , which can be explained by the
formation of a homogeneous lm. The presence of this
lm leads to a reduction in the double-layer capacitance
Cdl of the interface from the bare electrode value of 274

to 116 lF cm2 . Imidazoline is a heterocyclic compound


with three carbon and two nitrogen atoms in the ring.
Theoretical calculations were performed for both the
neutral and protonated forms, in order to give further
insight into the experimental results. In the 0.5 M HCl
solution that was used in this work, all molecules would
be protonated or even diprotonated [29]. The optimized
geometry of the molecules using the B3LYP hybrid
functional and the 6-311+G** basis set are presented in
Fig. 7 for amide, imidazoline and imidazolidine, and its
more stable protonated forms respectively. These results
are of predictive character with respect to the geometry
and reactivity of these compounds.
The bond distance, bond angles and dihedral angles
are shown in Table 2. It can be seen that the distance
N5C1 is smaller for imidazoline with respect to imidazolidine because to the presence of the double iminic

Fig. 7. Optimized B3LYP/6-311+G** structures of amide, imidazoline and imidazolidine and its more stable protonated forms. The NH hydrogen
bridge is also indicated. (For colours see online version of this article.)

J. Cruz et al. / Journal of Electroanalytical Chemistry 566 (2004) 111121

119

Table 2
 bond angles and dihedral in
Geometric parameters (bond length in A,
degrees) of amide, imidazoline and imidazolidine
Bond

Amide

Bond length/
A
C1N2
N2C3
C3C4
C4N5
N5C1
N5C6
C6C7
C7C8
Bond angle/
C1N2C3
N2C3C4
C3C4N5
C4N5C1
C4N5C6

Imidazoline

1.36
1.45
1.53
1.45

1.28
1.47
1.54
1.47
1.40
1.45
1.50
1.46

1.46
1.52
1.46
122.1
113.8
110

106.3
105.9
101.2
105.6
118.2

113.8

Dihedral angle/
C1N2C3C4
N2C1N5C4

94.4

12
13.4

Imidazolidine
1.47
1.48
1.56
1.48
1.50
1.47
1.54
1.48
103.7
106.5
104.1
105.7
113.7
26.7
36.27

bond N@C in the former. The bond and dihedral angles


show that the structure of the imidazolidine ring has an
enveloped conformation, and thus it is susceptible to
undergoing ring inversion. This suggests that the free
electron pair of nitrogen changes place and for this
reason, it is statistically more dicult for a reactant
complex to be obtained by means of a collision, thus
xing the molecule on the metal surface. In contrast, the
imidazoline ring, being plane, cannot present a conformational equilibrium because the position of the nonbonding electrons is less likely to change, thus helping
the interaction of imidazoline with the metal surface.
The ground state geometry of the inhibitor as well as
the nature of its frontier molecular orbitals, namely, the
HOMO and LUMO is involved in the activity properties
of the inhibitors. The natural analysis of charges is
shown in Table 3. These values indicate that heteroatoms have a nucleophilic character. The most favorable
sites for the interaction with the metal surface are the
following nitrogen atoms: N5 and N8 for amide, N2, N5
and N8 for imidazoline and N2 and N8 for imidazolidine, because they have a larger charge density. However, when the frontier molecular orbitals are analyzed,
Fig. 8, it is observed that for amide, the highest occupied
Table 3
Charge distribution for amide, imidazoline and imidazolidine
Atom

N2
N5
N8
O

Charge distribution
Amide

Imidazoline

Imidazolidine

)0.640
)0.684
)0.845
)0.644

)0.528
)0.553
)0.837

)0.668
)0.595
)0.847

Fig. 8. Molecular orbital HOMO for the amide, imidazoline and imidazolidine. (For colours see online version of this article.)

molecular orbital (HOMO) is in the nitrogen N5, this


being the favorite site for a nucleophilic attack. For the
imidazoline, on the other hand, the HOMO is localized
over the nitrogen atoms N2 and N5 of the plane ring;
consequently this is the preferred zone of the molecule
for interaction with the metal surface. Both the nitrogen
atoms can coordinate with carbon steel. Finally, the
amide presents a linear structure in which the hydrogen
atoms block the interaction of the most reactive sites
with the metal surface. This can explain the weak interaction with the metal. These results agree with reports
of the coordination compounds of metals with imidazolines [30,31]. The HOMO for imidazolidine is over nitrogen N5, which is the most reactive site of this
molecule but presenting a higher steric eect; this could
explain its almost null interaction with the metal surface.
To obtain more information about the local chemical
reactivity of these molecules, a study of Fukui indices
was carried out. These indices were calculated from the
optimized geometry of the neutral molecules by means
of the subtraction of one electron, N  1. The results are
shown in Table 4. Here, it is observed that for amide the
Fukui indices of the nitrogen N5 is more reactive than
N8, in agreement with its HOMO obtained value. The
Fukui indices for nitrogen N2 and N5 of imidazoline are
very similar. This seems to indicate that both atoms are

120

J. Cruz et al. / Journal of Electroanalytical Chemistry 566 (2004) 111121

Table 4
Calculated Fukui indices (nucleophilic)
Atom

N2
N5
N8
O

Fukui indices
Amide

Imidazoline

Imidazolidine

0.009
0.358
0.290
)0.038

0.252
0.271
0.090

0.035
0.477
0.084

involved in the chemical reactivity of this molecule with


the metal surface, while the N8 atom seems not to
participate in the reactivity, because its Fukui index was
only of +0.09. These results also agree with the HOMO
of this molecule (Fig. 8). Finally, imidazolidine showed
only one reaction site, localized on the nitrogen atom
N5, with a Fukui index of +0.477, but this atom presents
a higher steric eect, as is observed in the optimized
molecule (Fig. 7).
All possible protonated forms from the three organic
compounds studied, amide, imidazoline and imidazolidine, have been calculated, the more stable being the
molecules shown in Fig. 7. It can be noted from this
gure that the molecular structure of the protonated
form of imidazoline is very similar to the neutral form.
For instance, the bond-length C1N2 takes values of
 for the neutral form against 1.32 A
 for proton1.28 A
ated imidazoline. With respect to the bond-length C1

N5, its value changes from 1.47 for the neutral to 1.32 A
for the protonated form. These values seem to indicate
that a delocalization of p-electrons can take place in the
N2C1N5 bond. This means that the N2C1N5 bond
is of pp conjugation action between the p-orbital of the
N2 atom and p-MO of the C1N5 bond. The theoretical
calculations show that the molecular orbital HOMO of
the protonated form of imidazoline can be localized on
the N2C1N5 system. Owing to delocalization, the pelectron is easily translated to an Fe atom and is more
favorable for chemical adsorption of the N atom on the
metal surface.
Moreover, the imidazoline ring lies parallel to the
electrode, in a planar form, which means that the other
bond lengths, as well as angles, remain unchanged.
DFT analysis of protonated imidazoline shows that
nitrogen atoms have charge density values between
)0.45 and )0.58, showing that these atoms can play an
important role in the adsorption process of the molecule
with the metal surface, due to the fact that the molecules
exhibit a high electronic density and thus form organic
cations in acid environments.
The molecular structure of imidazolidine remains
unchanged, whereas for the case of amide, the formation
of an intramolecular hydrogen bridge N5HO confers
to the structure a greater degree of stability.
The molecular orbital for the amide is demonstrated
to be very important, because it is mainly localized over

the NAC@O system. Thus this site is more favorable for


the interaction between the molecule and the metal
surface. However, the reactivity of amide should be
chemically lower than that for the imidazoline, due to
steric factors. The linear structure of amide, Fig. 7, make
it possible that hydrogen atoms can block the interaction of the active sites of the molecule with the metal
surface.
The molecular structure of protonated imidazolidine
remains unchanged with respect to its neutral form, the
two N atoms on the ring remaining strongly blocked.
The only site for interaction could be the tertiary N
atom, but as has been demonstrated previously with the
Fukui analysis of the neutral compound, this atom does
not participate in the chemical reactivity.
From the theoretical calculations carried out with
both neutral and protonated forms of amide, imidazoline and imidazolidine, it can be concluded that the
structural and electronic properties of these molecules
are very similar. Consequently, theoretical calculations
with only the neutral form of these molecules could be
a good approach for the study of their potential
inhibition capability for metal corrosion in acidic
environments.

5. Conclusions
The relative inhibition eciencies of the three compounds studied, including their protonated forms, in a
deaerated 0.5 M HCl solution is found to be in the
following order: imidazoline  amide  imidazolidine.
With respect to imidazolidine, this compound showed a
very low corrosion inhibition eciency at high concentrations, being able to promote the corrosion even at
low concentrations.
The electrochemical techniques of EIS and Tafel
extrapolation provided consistent results during the
evaluation.
The EIS data provided additional information with
regard to the protective nature of the inhibitor lm.
The reasons for which imidazoline was the better
corrosion inhibitor are, among others, the plane geometry of the heterocyclic ring, thus favoring adsorption
through the double iminic bond N@C. The iminic nitrogen seems to be directly involved in the coordination
of the cycle with the metal surface.
The experimental results could be corroborated with
that obtained by means of the rst principles theoretical
calculations using the computational methodologies of
quantum chemistry, showing an electronic density more
important on the iminic nitrogen N2 in the imidazoline,
with respect to the analogous nitrogen in amide and
imidazolidine. The inhibitor eciencies of these compounds seem to be determined by the conjugation of
electronic factors and molecular geometry.

J. Cruz et al. / Journal of Electroanalytical Chemistry 566 (2004) 111121

Acknowledgements
J. Cruz, R. Martnez and E. Garca thank S. Cruz for
his observations and the Molecular Engineering Program from the Instituto Mexicano del Petr
oleo support
for this research, as well as the access to the supercomputer Silicon Graphics Origin 2000/48 of IMP. J.
Cruz gratefully acknowledges a grant from CONACYT.
The authors also wish to thank Pedro Rebollar for his
technical assistance.

References
[1] F. Bentiss, M. Lagrenee, M. Traisnel, Corrosion 56 (2000) 733.
[2] V.S. Sastry, Corrosion Inhibitors. Principles and Applications,
John Wiley & Sons, New York, 1998.
[3] M. Bouayed, H. Rabaa, A. Srhiri, J.-Y. Saillard, A. Ben Bachir,
A. Le Beuzed, Corros. Sci. 41 (1999) 501.
[4] F. Bentiss, M. Lagrenee, M. Traisnel, J.C. Hornez, Corros. Sci. 41
(1999) 789.
[5] F. Bentiss, M. Traisnel, M. Lagrenee, J. Appl. Electrochem. 31
(2001) 41.
[6] V. Jovancicevic, S. Ramachandran, P. Prince, Corrosion 55 (1999)
449.
[7] G.H. Awad, A.N. Assad, A.M. Abdel Gaber, S.S. Massoud,
Protect. Met. 33 (1997) 565.
[8] A.G. Gad AllaH, H. Moustafa, J. Appl. Electrochem. 22 (1992) 644.
[9] W. Urnie, R. De Marco, A. Jeerson, B. Kisella, J. Electrochem.
Soc. 146 (1999) 1751.
[10] S. Ramachandran, B. Tsai, M. Blanco, H. Chen, Y. Tang, W.A.
Goddard III, Langmuir 12 (1996) 6419.
[11] D. Wang, S. Li, Y. Ying, M. Wang, H. Xiao, Z. Chen, Corros.
Sci. 4 (1999) 1911.
[12] J. Cruz, L.M.R. Martnez-AguIlera, R. Salcedo, M. Castro, Int. J.
Quant. Chem. 85 (2001) 546.
[13] J.L. Riebsomer, J. Am. Chem. Soc. 70 (1948) 1679.

121

[14] R.N. Butler, C.B. ORegan, J. Chem. Soc. Perkin I (1986) 386.
[15] B.S. Kolomiets, L.P. Kuryaninova, V.V. Suchkov, Zh. Prikl.
Khim. 51 (1978) 668.
[16] S. Tang, Z. Jiang, Q. Mao, Oil & Gas (1994) 68.
[17] R.J. Ferm, J.L. Riebsomer, Chem. Rev. 54 (1954) 533.
[18] R.N. Butler, J.D. Thornton, P. Moynihan, J. Chem. Res. (S)
(1981) 84.
[19] J.A. Martin, F.W. Valone, Corrosion 41 (1985) 281.
[20] C.M. Blair, Corrosion 41 (1985) 616.
[21] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A.
Robb, J.R. Cheeseman, V.G. Zakrzewski, J.A. Montgomery Jr.,
R.E. Stratmann, J.C. Burant, S. Dapprich, J.M. Millam, A.D.
Daniels, K.N. Kudin, M.C. Strain, O. Farkas, J. Tomasi, V.
Barone, M. Cossi, R. Cammi, B. Mennucci, C. Pomelli, C.
Adamo, S. Cliord, J. Ochterski, G.A. Petersson, P.Y. Ayala, Q.
Cui, K. Morokuma, D.K. Malick, A.D. Rabuck, K. Raghavachari, J.B. Foresman, J. Cioslowski, J.V. Ortiz, A.G. Baboul, B.B.
Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R.
Gomperts, R.L. Martin, D.J. Fox, T. Keith, M.A. Al-Laham,
C.Y. Peng, A. Nanayakkara, C. Gonzalez, M. Challacombe,
P.M.W. Gill, B. Johnson, W. Chen, M.W. Wong, J.L. Andres, C.
Gonzalez, M. Head-Gordon, E.S. Replogle, J.A. Pople, Gaussian
98 (Revision A.7), Gaussian, Inc., Pittsburgh, PA, 1998.
[22] C. Lee, W. Yang, R.G. Parr, Phys. Rev. B 37 (1988) 785.
[23] A.D. Becke, J. Chem. Phys. 98 (1993) 5648.
[24] R.G. Parr, W.J. Yang, J. Am. Chem. Soc. 106 (1984) 4049.
[25] F. Bentiss, M. Lagrenee, M. Traisnel, J.C. Hornez, Corros. Sci. 41
(1999) 789.
[26] F. Bentiss, M. Traisnel, M. Lagrenee, J. Appl. Electrochem. 31
(2001) 41.
[27] L. Elkadi, B. Mernari, M. Traisnel, F. Bentiss, M. Lagrenee,
Corros. Sci. 42 (2000) 703.
[28] R. Cottis, S. Turgoose, Electrochemical Impedance and Noise,
NACE International, Houston, 1999, p. 36.
[29] D.A. Guzonas, D.E. Irish, G.F. Atkinson, Langmuir 5 (1988)
787.
[30] S. Ramachandran, B.L. Tsai, M. Blanco, H. Cheng, Y. Tang,
W.A. Goddard III, J. Phys. Chem. A 101 (1997) 83.
[31] K.F. Purcell, Inorganic Chemistry, Saunders Inc., Philadelphia,
1997.

Vous aimerez peut-être aussi